Molecular characterization and gene expression modulation of the alternative oxidase in a scuticociliate parasite by hypoxia and mitochondrial ...

Page created by Dwayne Chang
 
CONTINUE READING
Molecular characterization and gene expression modulation of the alternative oxidase in a scuticociliate parasite by hypoxia and mitochondrial ...
www.nature.com/scientificreports

            OPEN            Molecular characterization
                            and gene expression modulation
                            of the alternative oxidase
                            in a scuticociliate parasite
                            by hypoxia and mitochondrial
                            respiration inhibitors
                            Iría Folgueira1, Jesús Lamas2,3, Rosa Ana Sueiro1 & José Manuel Leiro1,3*

                            Philasterides dicentrarchi is a marine benthic microaerophilic scuticociliate and an opportunistic
                            endoparasite that can infect and cause high mortalities in cultured turbot (Scophthalmus maximus). In
                            addition to a cytochrome pathway (CP), the ciliate can use a cyanide-insensitive respiratory pathway,
                            which indicates the existence of an alternative oxidase (AOX) in the mitochondrion. Although AOX
                            activity has been described in P. dicentrarchi, based on functional assay results, genetic evidence
                            of the presence of AOX in the ciliate has not previously been reported. In this study, we conducted
                            genomic and transcriptomic analysis of the ciliate and identified the AOX gene and its corresponding
                            mRNA. The AOX gene (size 1,106 bp) contains four exons and three introns that generate an open
                            reading frame of 915 bp and a protein with a predicted molecular weight of 35.6 kDa. The amino
                            acid (aa) sequence of the AOX includes an import signal peptide targeting the mitochondria and the
                            protein is associated with the inner membrane of the mitochondria. Bioinformatic analysis predicted
                            that the peptide is a homodimeric glycoprotein, although monomeric forms may also appear under
                            native conditions, with EXXH motifs associated with the diiron active centers. The aa sequences of the
                            AOX of different P. dicentrarchi isolates are highly conserved and phylogenetically closely related to
                            AOXs of other ciliate species, especially scuticociliates. AOX expression increased significantly during
                            infection in the host and after the addition of CP inhibitors. This confirms the important physiological
                            roles of AOX in respiration under conditions of low levels of ­O2 and in protecting against oxidative
                            stress generated during infection in the host.

                             Alternative oxidase (AOX), also known as ubiquinol-oxidoreductase, is included within the diiron carboxylate
                             superfamily. The enzyme is characterized by cyanide resistance and is associated with maintenance of metabolic
                             homeostasis via mitochondrial electron transport in plants subjected to various stress ­conditions1,2. In plants,
                             AOX is considered a ubiquitous enzyme; however, in ­algae3, ­fungi4–9, ­protozoa10–13 and several invertebrate
                            ­species14 it is quite sporadic. AOX is involved in the mitochondrial respiratory chain and provides an alternative
                             pathway for the direct transfer of electrons from ubiquinone to oxygen, thus forming w    ­ ater15. It facilitates mito-
                             chondrial respiratory activity and therefore the functioning of the Krebs cycle after inhibition of the cytochrome
                             route due to the presence of nitric oxide, cyanide (both inhibitors of the complex IV) or antimycin A (complex
                             III inhibitor)1,12,16. However, AOX reduces the energy efficiency of respiration as only 1/3 of the ATP is produced

                            1
                             Department of Microbiology and Parasitology, Laboratory of Parasitology, Research Institute on Chemical
                            and Biological Analysis, University of Santiago de Compostela, Campus Vida, 15782 Santiago de Compostela,
                            Spain. 2Department of Fundamental Biology, Institute of Aquaculture, University of Santiago de Compostela,
                            Campus Vida, 15782 Santiago de Compostela, Spain. 3These authors contributed equally: Jesús Lamas and José
                            Manuel Leiro. *email: josemanuel.leiro@usc.es

Scientific Reports |   (2020) 10:11880                 | https://doi.org/10.1038/s41598-020-68791-9                                                1

                                                                                                                                              Vol.:(0123456789)
Molecular characterization and gene expression modulation of the alternative oxidase in a scuticociliate parasite by hypoxia and mitochondrial ...
www.nature.com/scientificreports/

                                      by pumping protons through complex ­I17. In plants, AOX participates in important biological processes, such
                                      as survival and adaptation to stress factors in biotic and abiotic environments, resistance to oxidative stress
                                      and maintenance of mitochondrial and cellular energy h         ­ omeostasis18,19. In parasitic protists, it is particularly
                                      associated with metabolic adaptation to the host and ­pathogenicity20–22. AOX is present in the mitochondria
                                      of protists such as the amoebozoans Acanthamoeba castellani23, A. polyphaga24 and Dictyostelium discoideum25,
                                      Trypanosoma brucei26, T. congolensi and T. evansi27, Plasmodium falciparum28, Cryptosporidium parvum29 and
                                      Blastocystis hominis16,22. AOX has also been detected in the Euglenozoa Euglena gracilis30, in the ciliates Parame-
                                      cium tetraurelia31, Tetrahymena pyriformis32 and Ichthyophthirius multifiliis33; and in the scuticociliate parasites
                                      Pseudocohnilembus persalinus34 and Philasterides dicentrachi13. In previous studies, we have demonstrated the
                                      presence of a cyanide-resistant alternative respiratory pathway (AP) in mitochondria of the scuticociliate parasite
                                      P. dicentrarchi, in addition to the conventional cytochrome pathway (CP)13. This provided physiological evidence
                                      for the presence of a cyanide-insensitive oxidase, which was recognised by antibodies raised against a synthetic
                                      peptide derived from a fully conserved C-terminal consensus motif from plant AOX ­isoforms13. In addition,
                                      AOX inhibitors such as salicylhydroxamic acid (SHAM) and 3,4,5-trihydroxybenzoic acid (propyl gallate, PG)
                                      significantly inhibit in vitro growth of the ­ciliate13,35,36.
                                          Although there is clear physiological evidence for the existence of an AOX in P. dicentrarchi, this enzyme has
                                      not previously been identified or characterized in detail at the molecular level. In the present study, we character-
                                      ized the P. dicentrarchi AOX gene and detected the AOX protein in the mitochondria. We also investigated the
                                      transcriptional modulation of this enzyme in vitro, in response to experimentally generated hypoxia, as well as
                                      in response to the presence of inhibitors of CP and AP mitochondrial respiration.

                                      Materials and methods
                                      Experimental animals and ethics statement.                Turbot (Scophthalmus maximus). Juvenile fish (50 g)
                                      were obtained from a fish farm in Galicia (NW Spain). The fish were held in 200 L fibreglass tanks with a water
                                      recirculation system, at 16 °C with constant aeration and a photoperiod of 12L:12D, and were fed daily with
                                      commercial pellets (Skretting, Burgos, Spain). The fish were acclimatized to the aquarium conditions for 2 weeks
                                      before the start of the experiments.
                                          Mice. Female CD-1 mice (Charles River Laboratories, USA) of weight 25 g were supplied by the Central
                                      Animal Husbandry Unit of the University of Santiago de Compostela (Spain).
                                          All animal experiments and protocols were conducted in accordance with Spanish and European Legislation
                                      (R.D. 53/2013 and Council Directive 2010/63/EU) and approved by the Institutional Animal Care and Use Com-
                                      mittee of the University of Santiago de Compostela (Spain). When required during all experiments, the animals
                                      were anaesthetized with isoflurane (mice) or 100 mg/L of tricaine methane sulphonate-MSS-222-(turbot) and
                                      finally euthanized by decapitation (mice) or by overdose of anaesthesia (turbot).

                                      Cell culture and experimental infections. The isolate of P. dicentrarchi37 used (I1) was obtained from
                                      ascites of experimentally infected turbot (Scophthalmus maximus)38. The isolate was grown axenically at 21 °C
                                      in complete L-15 ­medium39 under (a) normoxic conditions, in culture flasks provided with vented caps to allow
                                      aeration of the culture medium, or (b) under hypoxic conditions, by bubbling the culture flasks for 10 min with
                                      argon gas and tightly closing the flask with non-vented caps, after addition or removal of the c­ iliates13.

                                      Sequencing of the genome and transcriptome. For analysis of the P. dicentrarchi genome and tran-
                                      scriptome, trophonts ­(107) were concentrated by centrifugation, frozen in liquid nitrogen and sent on dry ice to
                                      Future Genomic Technologies (Leiden, Netherlands). For sequencing of the complete genome of the ciliate, a
                                      combination of short reading sequencing (Illumina technology) and long reading sequencing (Nanopore tech-
                                      nology) was used (Oxford Nanopore Technologies). For de novo assembly of the parasite genome, the data sets
                                      were combined using the TULIP program v0.440. Transcript sequences from Illumina RNA- Seq data (fragments
                                      of approximately 100 bp), obtained by amplification by SBS, were assembled using Trinity software (v2.6.5)41,
                                      included within the Galaxy application (https​://usega​laxy.org/). The assembled sequences were analyzed by
                                      homology, with Blastgo 5.0 software (Biobam, Spain), and annotated. The sequences that encode proteins that
                                      are potentially related to the ciliate AOX were then selected from the Tetrahymena thermophila gene and protein
                                      sequences database by using the BLASTx tool Wiki TGD (https​://www.cilia​te.org/blast​/blast​_link.cgi).

                                      Production of recombinant AOX (rAOX) in yeast cells. The complete nucleotide sequence that
                                      encodes P. dicentrarchi AOX was obtained with an open reading frame (ORF) search tool (ORF Finder; https​
                                      ://www.ncbi.nlm.nih.gov/orffi​nder/), from the annotated data obtained by analysis of a RNA-Seq experiment.
                                      The codons of the original nucleotide sequence in P. dicentrarchi were optimized using the Integrate DNA Tech-
                                      nologies (IDT) bioinformatics tool (https​://eu.idtdn​a.com/Codon​Opt), to produce a recombinant protein in
                                      the yeast Klyuveromyces lactis. The AOX gene nucleotide sequence was synthesized by Invitro GeneArt Gene
                                      Synthesis (ThermoFisher Scientific), amplified by PCR with the primers FaoxKl / RaoxKl: 5′-CGC CTC GAG
                                      AAA AGA ATG CAG TCC TTC GCC AGG AAA-3′/5′-ATA AGA ATG CGG CCG C TTA GTG GTG GTG
                                      ATG ATG GTG ATG ATG ATG ATG CTG CTC GGG GGA GTA AGG GTT 3′ and cloned in the integrative
                                      expression vector pKLAC2. The K. lactis Protein Expression kit (New England Biolabs, UK) was used to express
                                      the recombinant protein in yeast according to the manufacturer’s instructions and the protein secretion strategy,
                                      as previously ­described42.

                                      Obtaining anti‑rAOX immune serum. The anti-rAOX antiserum used in the immunoassays was
                                      obtained from CD-1 mice immunized intraperitoneally (i.p.) with 200 µg of rAOX. The antiserum was purified

          Scientific Reports |   (2020) 10:11880 |                 https://doi.org/10.1038/s41598-020-68791-9                                                  2

Vol:.(1234567890)
Molecular characterization and gene expression modulation of the alternative oxidase in a scuticociliate parasite by hypoxia and mitochondrial ...
www.nature.com/scientificreports/

                            by immobilized metal affinity chromatography (IMAC) with Ni-Sepharose42 and adsorbed for 30 min at room
                            temperature in 200 µL of a 1% chitosan hydrogel a­ djuvant42 (CH) (w/v) in PBS buffer, pH 7.0. At intervals of 15
                            and 30 days after the first immunization, mice were reimmunized with the same dose of rAOX and adjuvant.
                            Seven days after the second immunization, the mice were bled by decapitation and the blood thus obtained could
                            clot overnight at 4° C. The serum was clarified by centrifugation at 2000 × g for 10 min, mixed 1: 1 (v/v) with
                            glycerol, and stored at −20 °C until use.

                            Sodium dodecyl sulphate polyacrylamide gel electrophoresis (SDS‑PAGE) and Western‑blot-
                            ting. SDS-PAGE of the rAOX protein and a ciliate lysate (CL)—obtained from cultures maintained under
                            normoxic or hypoxic ­conditions44—and treated under reducing conditions (after the addition of 0.02 M dith-
                            iothreitol-DTT-) or under non-reducing (without the addition of DTT), was performed on linear 12.5% poly-
                            acrylamide minigels in a Mini-Protean Tetra cell system (BioRad, USA), as previously d­ escribed35. Once electro-
                            phoresis was completed, the gels were stained with a solution of GelCode Blue Stain Reagent (Thermo Scientific)
                            according to the manufacturer’s instructions.
                                Samples separated by electrophoresis were also analyzed by Western blot, with a slightly modified version of
                            a previously described ­protocol35. The samples were first incubated with serum from mice immunized with AOX
                            (anti-rAOX serum) and then with a polyclonal peroxidase-conjugated rabbit anti-mouse antibody (Dakopatts,
                            Denmark) at 1:800 dilution. The blots were stained by adding a chromogenic enzyme substrate solution con-
                            sisting of 0.003% ­H2O2 and 0.06% 3,30-diaminobenzidine tetrahydrochloride with 0.03% NiCl2 (DAB/NiCl2,
                            Sigma, USA).

                            Indirect immunofluorescence (IIF). For AOX immunolocalization in trophonts, an IIF immunoassay
                            was performed, as previously d­ escribed46. Briefly, ciliates were fixed in a solution of 4% formaldehyde in PBS,
                            permeabilized in a solution containing 0.3% Triton X-100 in PBS and blocked with a solution of 1% BSA. The
                            ciliates were then incubated with a mouse antiserum containing anti-rAOX antibodies diluted 1:100 in PBS
                            and with a secondary polyclonal rabbit anti-mouse immunoglobulin conjugated with fluorescein isothiocyanate
                            (FITC; DAKO, Denmark) and diluted 1:1,000. The ciliates were then visualized by fluorescence microscopy
                            (Zeiss Axioplan, Germany).

                            Transmission electron microscopy (TEM). For TEM analysis we followed the technique described by
                            Paramá et al.47. Briefly, the cultured ciliates were fixed in 2.5% (v/v) glutaraldehyde in 0.1 M cacodylate buffer
                            at pH 7.2, post-fixed in 1% (w/v) OsO4, pre-stained in saturated aqueous uranyl acetate, dehydrated in acetone
                            series and embedded in Spurr’s resin. Ultrathin sections were stained with 2.5% uranyl acetate in 50% ethanol
                            and 0.5% of lead citrate in 4.5 mM NaOH and viewed in a Jeol JEM-1011 transmission electron microscope
                            (Jeol, Japan) at an accelerating voltage of 100 kV.

                            In vitro growth assay. To determine the growth of the ciliates under conditions of normoxia and hypoxia,
                            we followed the protocol described by Mallo et al.13,35. Briefly, the ciliates were cultured at an initial concentra-
                            tion of 5 × 104 trophonts/mL in complete L-15 medium in culture flasks fitted with vented caps that allowed
                            aeration of the culture medium (normoxic conditions) and in tightly closed bottles bubbled with argon (hypoxic
                            conditions). To ensure hypoxic conditions, the L-15 medium was bubbled with argon for 10 min until gas satura-
                            tion, to completely displace the dissolved oxygen in the medium. The ciliates were then added, and the bottles
                            were hermetically sealed to prevent entry of oxygen. To determine growth of the cultures under conditions of
                            normoxia and hypoxia, aliquots of 100 μL were collected daily for 6 days and the number of ciliates was quanti-
                            fied in a Neubauer counting c­ hamber48. After collection of each sample to quantify the ciliates, the argon bub-
                            bling was repeated in all cases under the same conditions as described above to maintain the hypoxic conditions.

                             Reverse transcriptase‑quantitative polymerase chain reaction (RT‑qPCR). The RT-qPCR tech-
                            nique was basically performed as previously described, with slight ­variations35. Briefly, the total RNA extracted
                             from ­106 trophozoites/mL of P. dicentrarchi cultured under hypoxic/normoxic conditions, in the presence of
                             1 mM SHAM and 1 mM KCN (inhibitors of AOX and cytochrome-complex IV-, respectively), was isolated with
                             a NucleoSpin RNA isolation kit (Macherey–Nagel) according to the manufacturer’s instructions. The quality,
                             purity and concentration of RNA were quantified in a NanoDrop ND-1000 Spectrophotometer (NanoDrop
                             Technologies, USA). Two μg of sample RNA, 1·25 μM random hexamer primers (Promega) and 200 U of Molo-
                             ney murine leukemia virus reverse transcriptase (MMLV; Promega) were used for cDNA synthesis (RT). Quan-
                             titative polymerase chain reaction (qPCR) analysis was performed with AOX gene-specific primers forward/
                             reverse pair FqAOX/RqAOX (5′ TCG GAG ACT CCT TCG CTT AC-3′/5′-CCA TCC TTG GTC TCT TTC
                             CA-3′) and P. dicentrarchi elongation factor 1-alpha gene (EF-1α) (GenBank accession KF952262) forward/
                             reverse primer pair (FEF1A/REF1A: 5′-TCG CTC CTT CTT GCA TCG TT-3′/5′-TCT GGC TGG GTC GTT
                             TTT GT-3′) as housekeeping gene. The reaction mixture (10 μL) used in the qPCR contained 5 μL of PowerUp
                             SYBR Green Master Mix (Applied Biosystems), 300 nM of the primer pair, 1 μL of cDNA and RNase–DNase-
                             free water. The qPCR was run at 95 °C for 5 min, followed by 40 cycles at 95 °C for 10 s and 60 °C for 30 s, ending
                             with melting-curve analysis at 95 °C for 15 s, 55 °C for 15 s and 95 °C for 15 s. in a StepOnePlus Real Time Sys-
                             tem (Thermo Fisher Scientific). Finally, relative quantification of gene expression was determined by the 2­ −ΔΔct
                            ­method49 with the software included in the equipment.

Scientific Reports |   (2020) 10:11880 |               https://doi.org/10.1038/s41598-020-68791-9                                              3

                                                                                                                                          Vol.:(0123456789)
Molecular characterization and gene expression modulation of the alternative oxidase in a scuticociliate parasite by hypoxia and mitochondrial ...
www.nature.com/scientificreports/

                                      Figure 1.  (A) Schematic representation of the structure of the AOX gene of P. dicentrarchi, indicating the
                                      position of exons (squares) and introns (lines) in the nucleotide sequence, as well as their translation to
                                      mRNA. Linear scale = 100 nucleotides (B) Philasterides dicentrarchi AOX mRNA, complete cds (GenBank:
                                      MH427340.1), together with its corresponding translation into amino acids (GenBank: QAR17767.1),
                                      including the mitochondrial targeting sequences (MTS), the diiron binding motifs [ion binding site] (DBM),
                                      and predictions for O-ß-GlcNAc attachment sites. (C) Diagram of the bioinformatic prediction of the cellular
                                      location of the amino acid sequences of the AOX provided by the LocTree 3 program (Rostlab, Technical
                                      University of Munich, Germany; URL: https​://rostl​ab.org/servi​ces/loctr​ee3/).

                                      Bioinformatic and statistical analysis. Functional analysis of proteins and classification into different
                                      families to predict the domains and important sites was carried out with INTERPRO ­software50. The transmem-
                                       brane topology and location of signal peptide cleavage sites in amino acid (aa) sequences were predicted with the
                                       PHOBIUS ­program51. The N-terminal region of the protein that may contain a mitochondrial signal sequence
                                      and its cleavage site were analyzed with the MITOPROT II -v1.101 p       ­ rogram52. The physicochemical param-
                                       eters were predicted for a given protein with the The PROTPARAM ­tool53. The O-ß-GlcNAc attachment sites
                                       were predicted with the Y ­ INGOYANG54 and O     ­ GLCNACSCAN55 programs. For visualization of proteoforms
                                       and integration of the protein topology, the PROTTER bioinformatics tool was u      ­ sed56. Sub-cellular localiza-
                                       tion was predicted with the LOCTREE3 ­program57. The SWISS-MODEL protein server was used for model-
                                       ling protein structures and c­ omplexes58. The aa sequences were aligned using the CLUSTAL OMEGA multiple
                                      sequence alignment program 59, and the phylogenetic signal/noise ratio was improved with the TRIMAI tool
                                      for automated alignment ­trimming60. The phylogenetic trees were constructed by the Maximum Likelihood
                                      (ML) method, with the JTT m      ­ odel61. Branch support was given with 1,000 bootstrap r­ eplicates62 in MEGAX
                                      ­software63. The Bayesian inference (BI) analysis was performed with MRBAYES 3.2.664.
                                           The values shown in the text and figures are means ± SEM. One-way analysis of variance (ANOVA) was used
                                       for comparison of more than two samples, and the Tukey-Kramer test was used for pairwise comparisons. The
                                       Student’s t-test was used for comparison of two samples. In all cases, differences were considered significant at
                                       P < 0.05.

                                      Results
                                      Overall structure of AOX in P. dicentrarchi.           The complete sequences of the AOX gene and cDNA were
                                      deposited in the NCBI GenBank database, with accession numbers MN193569 and MH427340, respectively.
                                      The gene encoding the AOX has a total length of 1,107 bp and contains 4 exons and 3 introns, with an average
                                      intron size of around 60 bp (Fig. 1A). The coding sequence contains 918 bp that encodes a protein of 305 aa
                                      with an estimated molecular mass of 35,638.68 daltons and a theoretical pI of 9.03. Bioinformatic analysis also
                                      indicated that this protein lacks a signal peptide; however, the MitoPro II program predicted the existence of
                                      an import signal peptide targeting the mitochondria MQSFARKFCTSSALV with a probability of 0.9503 and a
                                      cleavage site in aa 16 (Fig. 1B), as also confirmed by the bioinformatic prediction of cellular localization in the

          Scientific Reports |   (2020) 10:11880 |               https://doi.org/10.1038/s41598-020-68791-9                                             4

Vol:.(1234567890)
www.nature.com/scientificreports/

                            Figure 2.  Structural modelling and active sites of P. dicentrarchi AOX. (A) Model of spatial structure for AOX
                            homodimeric complex showing regions where the active sites containing the diiron binding motifs in the two
                            polypeptide chains (1 and 2) of the enzyme are located (circles). (B) Diiron active site in chain 1, showing
                            diiron and hydroxo atoms as well as the interactions of the diiron complex with the glutamate and histidine
                            residues. (C) Diiron active site in chain 2, showing diiron and hydroxyl atoms as well as the interactions of the
                            diiron complex with the glutamate and histidine residues. In both cases, the diiron site (iron moieties in orange,
                            hydroxyl in red) is contained within a four alpha-helix bundle. The images of build model were generated by
                            the Swiss-Model program (BIOZENTRUM, The Centre for Molecular Life Sci, University of Basel, Switzerland;
                            URL: https​://swiss​model​.expas​y.org/).

                            mitochondria through the LocTree program (Fig. 1C). Bioinformatic analysis with the InterPro program (which
                            provides functional analysis of proteins by classifying them in families and predicts important domains and sites
                            in the sequence) showed that the protein belongs to the AOX family, i.e. that it has 6 diiron binding motifs and
                            is O-glycosylated in the aa S3, T10, S11, T55, S117, S118, S121, T215, T259, S302 (Fig. 1B). Prediction of the
                            biological function from the amino acid sequence of the AOX to establish the gene ontology (GO) terms indi-
                            cates that this protein is associated with oxidation–reduction processes and with a biological function involving
                            alternative oxidase activity. Analysis of the topology of the aa sequence using the Phobius bioinformatic program
                            revealed that this protein has two transmembrane regions associated with the mitochondrial inner membrane
                            between aa 125 and 146 and 188–211, a region located between aa 1–125 and between aa 212–305 that is located
                            inside the mitochondrial matrix, and a region located between aa 147 and 187 that is located towards the inter-
                            membrane space of the mitochondria. However, the program for predicting transmembrane helices in proteins
                            (TMHMM-ver. 2.0) only identified a single transmembrane region associated with the mitochondrial inner
                            membrane between aa 188–211, a region between aa 1–188 that is located inside the mitochondrial matrix, and
                            a region located between aa 212–305 directed towards the mitochondrial intermembrane space.

                            Structure of diiron active site of AOX. The protein structure modelling of AOX generated by the Swiss-
                            Model and Swiss-PBD Viewer programs showed that the protein is homodimeric (Fig. 2A) and that each mono-
                            mer contains an active site formed by two diiron centers and a OH ion ­(Fe3+–OH–Fe3+) (Fig. 2A,B). The compu-
                            tational ligand modeling indicated that Fe.1 has three ligand contacts with aa, two with glutamate (E133, E172)
                            and one with histidine H175) residues, and one with OH; while Fe.2 interacts with three glutamate residues
                            (E172, E223 and E279) and with OH group. Similarly, the active site of the other monomer also contains two
                            diiron centers that interact with an OH group and with the same glutamate and histidine residues located in that
                            monomer (Fig. 2C).

                            Phylogenetic analysis. We initially aligned the AOX amino acid sequences of three representative P.
                            dicentrarchi strains (I1, B1 and C1) in order to estimate the degree of identity (Supplementary information).
                            The AOX sequences in these P. dicentrarchi strains are highly conserved: there is a 100% identity between the
                            AOX sequences of strains I1 and C1, and 99.67% identity with strain B1, which corresponds to a single change
                            in a phenylalanine for a valine in aa 8 (Supplementary information). Analysis of the phylogenetic relationships
                            between the aa sequence of the P. dicentrarchi AOX and those of AOXs of other Ciliophora indicate that are
                            closely related to the AOX of the scuticociliate species Pseudocohnilembus persalinus (Fig. 3). The AOXs of Cili-
                            ophora species were also found to be more closely related to the AOX of Metazoan and Fungi than to Viridiplan-
                            tae and other SAR species such as, for example, the Apicomplexa (Fig. 3).

                            Biochemical analysis and subcellular localization of P. dicentrarchi AOX. The biochemical prop-
                            erties of the AOX protein of P. dicentrarchi were investigated by SDS-PAGE and immunological techniques. To
                            perform the biochemical analysis, a recombinant AOX was generated and expressed in the yeast K. lactis. The
                            complete cDNA encoding the AOX was chemically synthesized and cloned in the yeast to produce the recombi-

Scientific Reports |   (2020) 10:11880 |               https://doi.org/10.1038/s41598-020-68791-9                                            5

                                                                                                                                        Vol.:(0123456789)
www.nature.com/scientificreports/

                                      Figure 3.  Phylogenetic analysis of amino acid sequences of P. dicentrarchi AOX inferred by Maximum
                                      Likelihood method (ML) and Bayesian inference (BI) methods, respectively. The analysis includes sequences
                                      of AOXs from representative species of Metazoa, Viridiplantae, Fungi and SAR. The GenBank access number
                                      is included after the name of the species. The tree with the highest log likelihood (− 8,244.09) is shown.
                                      The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1,000
                                      replicates) are shown beside the branches and represent the bootstrap value for ML analysis (black) and
                                      posterior probability value of BI analysis (red). Dashes (–) indicate disagreement between the ML and BI
                                      analysis. The tree is drawn to scale, with branch lengths in the same units as those of the evolutionary distances
                                      used to infer the phylogenetic tree. The evolutionary distances were computed using the Poisson correction
                                      method and are in the units of the number of amino acid substitutions per site. Analysis involved 28 amino acid
                                      sequences. All positions containing gaps and missing data were eliminated using the trimAl program. In total,
                                      226 positions were included in the final dataset. Evolutionary analyses were conducted with the MEGAX and
                                      MrBayes programs. The scale bar corresponds to 20 substitutions per 100 aa positions.

                                      nant rAOX protein, which was purified and analyzed by SDS-PAGE, verifying the size at about 35 kDa (Fig. 4B,
                                      lanes 1 and 2). Once it was confirmed that the protein is expressed in yeast, a mouse polyclonal antibody was
                                      generated for use in immunological analysis. Western blot analysis with the anti-rAOX antibody verified the
                                      existence of a single recognition band on the recombinant protein (Fig. 4A) under both reducing (lane 4) and
                                      non-reducing conditions (lane 3). However, addition to the antibody of a complete lysate of the ciliate separated
                                      under reducing conditions (Fig. 4A, lane 2) or non-reducing conditions (Fig. 4A, lane 1) led to the presence
                                      of a single 35 kDa band, in the first case, and of two bands of 35 kDa (poorly stained) and 70 kDa (intensely
                                      stained), in the second case. This indicates that the enzyme, may appear in a monomeric or dimeric state (in its
                                      native state), in the latter case, of size 75 kDa. To demonstrate the cellular location of the AOX, we performed
                                      an indirect immunofluorescence assay using the polyclonal anti-rAOX antibody (Fig. 4C). This antibody gener-
                                      ates intense labeling immediately below the plasma membrane at the location of the mitochondrion (Fig. 4D).

                                      Effect of hypoxia and mitochondrial respiration inhibitors on AOX expression. We first moni-
                                      tored the growth kinetics of P. dicentrarchi exposed to normoxia and hypoxia conditions. Growth of the ciliate
                                      was not altered under either condition (Fig. 5A), clearly indicating that the ciliate can grow normally under
                                      hypoxic conditions. We then analyzed AOX gene expression during different phases of growth under normoxia
                                      conditions, observing that the levels of transcripts produced under these circumstances were maintained at con-
                                      stant levels during the first 4 days of ciliate growth (i.e. during the logarithmic phase of growth). However, when
                                      ciliate growth reached the stationary phase, the levels AOX transcript amount increased significantly (Fig. 5B).
                                      The same was observed when the AOX protein expression was analyzed by Western blot with the polyclonal
                                      anti-rAOX antibody (Fig. 5C).

          Scientific Reports |   (2020) 10:11880 |               https://doi.org/10.1038/s41598-020-68791-9                                            6

Vol:.(1234567890)
www.nature.com/scientificreports/

                            Figure 4.  Biochemical characterization and cellular localization of the AOX of P. dicentrarchi. (A) Western
                            blot analysis with a polyclonal anti-recombinant AOX (anti-rAOX) on a total ciliate lysate (CL, lanes 1 and 2)
                            from trophonts of P. dicentrarchi maintained for 24 h in hypoxia and on the rAOX (lines 3 and 4). The proteins
                            were separated by SDS-PAGE under non-reducing (− DTT, lanes 1 and 3) and reducing (+ DTT, lanes 2 and
                            4) conditions. (B) SDS-PAGE profile of rAOX stained with Coomassie blue under non-reducing (lane 1)
                            and reducing conditions (lane 2). Arrows indicate the presence of recognition bands. MW molecular weight
                            markers. (C) Representative photomicrograph of an immunofluorescence assay performed on a trophont of P.
                            dicentrarchi and using a polyclonal anti-rAOX antibody in which intense fluorescent mitochondrial labelling
                            is observed (arrows). (D) TEM microphotograph of a trophont of P. dicentrarchi showing the location of the
                            mitochondria below the plasma membrane (arrows).

                               AOX gene expression during an infection in turbot was also analyzed. Turbot were experimentally infected
                            with ciliates from normoxic cultures in the logarithmic phase of growth, and expression of the AOX gene was
                            determined in the ciliates obtained from fish peritoneal cavity at 24 h post-infection (Fig. 5D). AOX expression
                            was higher in ciliates obtained from infected fish than in ciliates maintained in vitro under normoxia conditions
                            (Fig. 5D).
                               Finally, we carried out an experiment in which expression of the AOX gene was analyzed in ciliates maintained
                            under normoxia and hypoxia and treated with cyanide and SHAM, classical inhibitors of mitochondrial CP and
                            AP, respectively. The AOX transcript amount under hypoxia were significantly higher than under normoxia
                            (Fig. 5E). Similarly, in presence of SHAM, AOX gene expression was also higher in hypoxic conditions (Fig. 5E).
                            However, in presence of cyanide, the AOX expression levels were similar under both normoxia and hypoxia
                            (Fig. 5E). The gene expression results were similar to those obtained by Western blot, after analysis of the AOX
                            protein levels in the ciliates by using a polyclonal anti-rAOX antibody (Fig. 5F).

                            Discussion
                            AOX is a terminal ubiquinol oxidase that participates in an alternative mitochondrial pathway that couples
                            oxidation of ubiquinol to the reduction of ­O2 to ­H2O without contributing to the formation of an electrochemi-
                            cal ­gradient18. The first evidence for the existence of an AOX was based exclusively on physiological studies in
                            thermogenic plant tissues, in which mitochondrial respiration resistant to inhibitors of the classical cytochrome
                            pathway was attributed to the presence of cyanide-insensitive terminal o      ­ xidase22,65. The subsequent use of mito-
                            chondrial purification techniques enabled isolation and purification of the AOX protein and generation of
                            mono- and polyclonal antibodies, which led to the identification of AOX by immunochemical ­techniques4. In
                            recent years, molecular techniques have enabled identification of the genes that encode AOX and confirma-
                            tion of the existence of this enzyme in numerous organisms. These techniques have also enabled study of the
                            domains involved in enzymatic catalysis and of the topography of the mitochondrial membrane enzyme and its
                            post-translational regulation, as well as reconstruction of the evolution of the enzyme in different taxonomic
                            ­groups21. The genes encoding AOX have been found in all higher plants and in some algae, fungi, protists and
                            invertebrates, but not in v­ ertebrates14,15,22,66,67. We have found evidence of the existence of an alternative oxidase
                            pathway in the mitochondria of the fish parasite P. dicentrarchi, by using functional physiological techniques,
                            which measure oxygen consumption after the addition of inhibitors of the cytochrome pathway, and immuno-
                            chemical assays, which use polyclonal antibodies against a fully conserved C-terminal consensus motif from
                            isoforms 1 and 2 of plant ­AOX13. However, until now we did not have any information about the molecular
                            characteristics of this enzyme in P. dicentrarchi. After assembly of the genome and transcriptome of P. dicentrar-
                            chi by next-generation sequencing (NGS), we used bioinformatic tools to identify and predict the structure of
                            the gene and the mRNA that encodes the AOX. The AOX of P. dicentrarchi is encoded by a single nuclear gene
                            consisting of 4 exons and 3 introns. In many plant species, AOX enzymes are encoded by a multigene f­ amily68;
                            however, in fungi and in some protozoa the genes encoding AOX are less diverse than in plants and, in many
                            cases, this enzyme is encoded by a single nuclear g­ ene69. The typical structure of the AOX gene is characterized
                             by the presence of four exons separated by three gene-specific introns of different l­ engths70,71. After transcription

Scientific Reports |   (2020) 10:11880 |                https://doi.org/10.1038/s41598-020-68791-9                                                7

                                                                                                                                             Vol.:(0123456789)
www.nature.com/scientificreports/

                                      Figure 5.  Analysis of the expression of the P. dicentrarchi AOX gene. (A) Ciliate growth kinetics under
                                      conditions of normoxia and hypoxia. (B) AOX gene transcription levels in trophonts maintained under
                                      normoxia for 6 days. (C) Western blot analysis using a polyclonal anti-rAOX antibody and ciliated lysates
                                      obtained on different days from a normoxic culture. (D) mRNA levels in trophonts obtained from in vitro
                                      cultures under normoxia (control) conditions, and in the same ciliates after intraperitoneal injection in turbot
                                      and subsequent extraction at 24 h post-infection (24 h). (E) AOX transcript amount in trophonts maintained
                                      under normoxia and hypoxia for 6 days (Control, C) and treated with 1 mM of KCN or salicylhydroxamic
                                      acid (SHAM). (F) Western blot analysis with a polyclonal anti-rAOX antibody in ciliate lysate of trophonts
                                      maintained in the same conditions as in (E). The results presented in the graphs correspond to the mean
                                      values ± the standard error (SE; n = 5). Statistically significant difference between groups are indicated by
                                      different letters (P < 0.05).

                                       and splicing, the AOX gene in P. dicentrarchi generated a mature mRNA that encodes a protein of an estimated
                                       molecular weight of 35,638.68 daltons, a size like the standard sizes of the AOXs of various organisms ranging
                                       between 32–36 ­kDa15. AOX is an integral interfacial membrane protein that interacts with a single leaflet of the
                                       lipid ­bilayer26. The first studies of the membrane localization of AOX in plant mitochondria showed that AOX
                                         has two transmembrane helices and a short loop located in the m     ­ atrix72,73. However, later topological studies
                                        indicated that AOX is an integral monotopic membrane protein that embeds into a single face (inner) of the
                                      ­membrane74–76. There is some controversy regarding prediction of the transmembrane topology of the AOX in
                                       P. dicentrarchi. According to the Phobius program, AOX has two transmembrane helix regions, while other bio-
                                       informatic programs predict a single transmembrane helix region. In the first case, it is assumed that the C- and
                                       N-terminal domains of the protein will be located in the mitochondrial matrix; however, in the second model the
                                       N-terminal motif would be located outside the inner membrane and the C-terminal towards the mitochondrial
                                       ­matrix77. To resolve these differences, new computational methods are probably needed, as suggested by other
                                        ­authors78. In any case, until the crystal structure of AOX of P. dicentrarchi is available, it will not be possible to
                                       definitively clarify the topology of this enzyme.

          Scientific Reports |   (2020) 10:11880 |                https://doi.org/10.1038/s41598-020-68791-9                                                 8

Vol:.(1234567890)
www.nature.com/scientificreports/

                                  Analysis of AOX crystal structure indicates that it is a homodimer with a non-haem diiron carboxylate protein
                             in which the metal atoms are ligated to four fully conserved glutamate and histidine ­residues22, 79–81. Although the
                             AOX family contains highly conserved active sites, small differences have been observed in ligands associated
                             with the active center in several A ­ OXs22. P. dicentrarchi has the typical C-terminal E-X-X-H m       ­ otif73,82, in which
                             the conserved glutamate and histidine residues participate in the coordination of two iron a­ toms21.
                                  Phylogenetic studies of AOX sequences indicate that Viridiplantae, Metazoan, Fungi and SAR form mono-
                             phyletic ­clades25,83. On the other hand, Maximum Likelihood methods have revealed the existence of two main
                             clades of AOX proteins: a clade containing AOXs of Ascomycota and Basidiomycota fungi, invertebrates, kine-
                             toplastids and alveolate protozoa, and another clade that includes plants, algae and various fungal g­ roups83. In
                             our study, phylogenetic analysis by ML and BI methods also indicated the existence of two main clades of AOXs:
                             and Sar (apicomplexans and kinetoplastids), and another including Metazoa, Fungi and Sar (Ciliophora). The
                             present results also indicate that the AOX of P. dicentrarchi is closely related to AOXs of other ciliate species,
                             especially the scuticociliate P. persalinus, but it is relatively distant from other Sar members as apicomplexans and
                             kinetoplastids. In addition, although the enzyme sequence is highly conserved among strains of P. dicentrarchi,
                             some differences could be used, together with sequences of other proteins, to carry out multigenic genotyping,
                             to enable accurate, intraspecific characterization of the isolates of this s­ cuticociliate84.
                                  Plant studies indicate that AOX is an integral homodimeric membrane protein in which the monomers have
                             an estimated molecular mass of 40 kDa and a mitochondrial directed sequence that is cleaved to generate mature
                             forms with a molecular mass of over 32 ­kDa75. Immunochemical and immunohistochemical analysis indicates
                             that P. dicentrarchi have monomeric forms (of about 35 kDa) and homodimeric forms (of 70 kDa) associated
                             with the mitochondrial inner membrane.
                                  These results strongly suggest that a disulfide bond can link two monomers (there are some Cys residues that
                             could be involved, eg. Cyst77 of AOX sequence). This is what occurs in plants, where the status of this disulfide
                             bond impacts in enzyme activity, thus when there is not disulfide bond between two monomers (reduced form),
                             the AOX activity is strongly activated by allosteric effectors, while the oxidized form (disulfide bond between two
                             monomers) has little activity, even with high concentrations of the allosteric ­effectors79,85. Therefore, the results
                             obtained indicate that P. dicentrarchi may show two AOX forms: one monomeric of 35 kDa and other dimeric
                             of 70 kDa, due to the presence of a mixture of “reduced” and “oxidized” forms in the mitochondrial membrane.
                                  Phylogenetic analysis of AOX from different organisms has indicated the existence of a monophyletic clade
                             among benthic marine animals and the presence of AOX in these species, which is probably related to a mecha-
                             nism of adaptation to variations in oxygen concentration, hypoxic conditions and increased hydrogen sulphide
                            ­concentration86–88. This hypothesis can also be extended to P. dicentrarchi, a benthic (like most scuticociliates)
                             marine organism, which is also ­microaerophilic13,89.
                                  Expression of AOX has been reported to be modulated in response to hypoxia and anoxia or after
                             ­reoxygenation18. AOX has a lower affinity for ­O2 than cytochrome oxidase, possibly indicating higher expression/
                             activity under conditions of low oxygen ­concentration90,91. It has also been suggested that AOX expression can be
                             increased in plants under hypoxia to reduce the formation of superoxide ­(O2−), nitric oxide (NO) and peroxyni-
                             trite ­(ONOO-), as well as to reduce the toxic effect of ­H2O292–94. P. dicentrarchi is a microaerophilic microorgan-
                             ism capable of surviving in conditions of low oxygen c­ oncentrations13. Under these conditions, amount of AOX
                             transcript and of protein increase significantly in the ciliate. Gene transcription and protein expression increase
                             during the stationary phase of growth, during which O         ­ 2 levels in the culture are l­ow13. The low concentrations
                             of ­O2 and the increased levels of reactive oxygen species (ROS) are also associated with infection in the host. In
                             the first case, ­O2 levels during the endoparasitic phase are low, while ROS levels are high as a result of the innate
                             immune response produced during i­ nfection95,96. Both the infection and the innate humoral immune response
                             generated by the host induce a significant increase in AOX transcription in the ciliate.
                                  Agents that inhibit the cytochrome bc1 significantly induce AOX ­transcription97,98. Hypoxia, generated either
                             by the removal of O  ­ 2 from the culture medium or by the addition of cyanide, significantly induces the transcrip-
                             tion and expression of A   ­ OX13. On the other hand, the addition of the AOX inhibitor, S­ HAM99, only increases the
                             transcription of the enzyme under hypoxic conditions. The data of the increase in AOX transcription obtained
                             in this study are closely related to the results obtained at the physiological level. Thus, the determination of the
                             mitochondrial oxygen consumption shows that under hypoxia conditions and in the presence of inhibitors of
                             the cytochrome pathway, the ciliate preferably uses the alternative pathway ­respiration13.
                                  In conclusion, we have identified the gene and the mRNA that encode a protein belonging to the AOX family,
                              not previously characterized at the molecular level in the scuticociliate P. dicentrarchi. In this ciliate, the AOX
                              occurs in the monomeric and dimeric forms and is associated with the inner membrane of the mitochondria. The
                              amino acid sequence was found to be highly conserved among the different strains of this ciliate and is closely
                              related to AOXs of ciliate species and, particularly, to the AOX of scuticociliates. Gene transcription increased
                              significantly under conditions of hypoxia and after infection and addition of inhibitors of cytochrome pathway
                              respiration. The molecular characterization of AOX, together with the functional evidence of its role in adapting
                              to parasitism and in resistance to oxidative stress induced after infection, opens a promising route for developing
                              structure-based antiparasitic drug design strategies through molecular docking.

                            Data availability
                            The datasets generated during and/or analysed during the current study are available from the corresponding
                            author on reasonable request.

                            Received: 27 December 2019; Accepted: 1 July 2020

Scientific Reports |   (2020) 10:11880 |                 https://doi.org/10.1038/s41598-020-68791-9                                                     9

                                                                                                                                                  Vol.:(0123456789)
www.nature.com/scientificreports/

                                      References
                                       1. Moore, A. L. & Albury, M. S. Further insights into the structure of the alternative oxidase: From plants to parasites. Biochem. Soc.
                                          Trans. 36, 1022–1026 (2008).
                                       2. Albury, M. S., Elliott, C. & Moore, A. L. Towards a structural elucidation of the alternative oxidase in plants. Physiol. Plant. 137,
                                          316–327 (2009).
                                       3. Ünlü, E. S., Ünüvar, ÖC. & Aydın, M. Identification of alternative oxidase encoding genes in Caulerpa cylindracea by de novo
                                          RNA-Seq assembly analysis. Mar. Genom. 46, 41–48 (2019).
                                       4. Lambowitz, A. M., Sabourin, J. R., Bertrand, H., Nickels, R. & McIntosh, L. Immunological identification of the alternative oxidase
                                          of Neurospora crassa mitochondria. Mol. Cell. Biol. 9, 1362–1364 (1989).
                                       5. Li, Q. et al. Cloning and analysis of the alternative oxidase gene of Neurospora crassa. Genetics 142, 129–140 (1996).
                                       6. Veiga, A., Arrabaca, J. D. & Loureiro-Dias, M. C. Cyanide-resistant respiration, a very frequent metabolic pathway in yeasts. FEMS
                                          Yeast Res. 3, 239–245 (2003).
                                       7. Juarez, O. et al. The physiologic role of alternative oxidase in Ustilago maydis. FEBS J. 273, 4603–4615 (2006).
                                       8. Williams, B. A. et al. A broad distribution of the alternative oxidase in microsporidian parasites. PLoS Pathog. 6, e1000761 (2010).
                                       9. Luévano-Martínez, L. A. et al. Mitochondrial alternative oxidase is determinant for growth and sporulation in the early diverging
                                          fungus Blastocladiella emersoni. Fungal Biol. 123, 59–65 (2019).
                                      10. Jarmuszkiewicz, W., Wagner, A. M., Wagner, M. J. & Hryniewiecka, L. Immunological identification of the alternative oxidase of
                                          Acanthamoeba castellanii mitochondria. FEBS Lett. 411, 110–114 (1997).
                                      11. Castro-Guerrero, N. A., Krab, K. & Moreno-Sánchez, R. The alternative respiratory pathway of Euglena mitochondria. J. Bioenergy
                                          Biomembr. 36, 459–469 (2004).
                                      12. Chaudhuri, M., Ott, R. D. & Hill, G. C. Trypanosome alternative oxidase: From molecule to function. Trends Parasitol. 22, 484–491
                                          (2006).
                                      13. Mallo, N., Lamas, J. & Leiro, J. M. Evidence of an alternative oxidase pathway for mitochondrial respiration in the scuticociliate
                                          Philasterides dicentrarchi. Protist 164, 824–836 (2013).
                                      14. McDonald, A. E., Vanlerberghe, G. C. & Staples, J. F. Alternative oxidase in animals: Unique characteristics and taxonomic distri-
                                          bution. J. Exp. Biol. 212, 2627–2634 (2009).
                                      15. Araújo Castro, J., Gomes Ferreira, M. D., Santana Silva, R. J., Andrade, B. S. & Micheli, F. Alternative oxidase (AOX) constitutes
                                          a small family of proteins in Citrus clementina and Citrus sinensis L. Osb. PLoS ONE 12, e0176878 (2017).
                                      16. Moore, A. L. et al. Unraveling the heater: New insights into the structure of the alternative oxidase. Annu. Rev. Plant Biol. 64,
                                          637–663 (2013).
                                      17. Kumari, A., Pathak, P. K., Bulle, M., Igamberdiev, A. U. & Gupta, K. J. Alternative oxidase is an important player in the regulation
                                          of nitric oxide levels under normoxic and hypoxic conditions in plants. J. Exp. Bot. https​://doi.org/10.1093/jxb/erz16​0 (2019)
                                          (Epub ahead of print).
                                      18. Vanlerberghe, G. C. Alternative oxidase: A mitochondrial respiratory pathway to maintain metabolic and signaling homeostasis
                                          during abiotic and biotic stress in plants. Int. J. Mol. Sci. 14, 6805–6847 (2013).
                                      19. Pu, X. et al. Roles of mitochondrial energy dissipation systems in plant development and acclimation to stress. Ann Bot. 116,
                                          583–600 (2015).
                                      20. Walker, R. Jr., Saha, L., Hill, G. C. & Chaudhuri, M. The effect of over-expression of the alternative oxidase in the procyclic forms
                                          of Trypanosoma brucei. Mol. Biochem. Parasitol. 139, 153–162 (2005).
                                      21. Rogov, A. G., Sukhanova, E. I., Uralskaya, L. A., Aliverdieva, D. A. & Zvyagilskaya, R. A. Alternatiive oxidase: Distribution, induc-
                                          tion, properties, structure, regulation, and functions. Biochemistry (Moscow) 79, 1615–1634 (2014).
                                      22. May, B., Young, L. & Moore, A. L. Structural insights into the alternative oxidases: Are all oxidases made equal?. Biochem. Soc.
                                          Trans. 45, 731–740 (2017).
                                      23. Antos-Krzeminska, N. & Jarmuszkiewicz, W. Functional expression of the Acanthamoeba castellanii alternative oxidase in Escheri-
                                          chia coli; regulation of the activity and evidence for AcAox gene function. Biochem. Cell Biol. 92, 235–241 (2014).
                                      24. Henriquez, F. L. et al. Acanthamoeba alternative oxidase genes: Identification, characterization and potential as antimicrobial
                                          targets. Int. J. Parasitol. 39, 1417–1424 (2009).
                                      25. Kimura, K., Kuwayama, H., Amagai, A. & Maeda, Y. Developmental significance of cyanide-resistant respiration under stressed
                                          conditions: Experiments in Dictyostelium cells. Dev. Growth Differ. 52, 645–656 (2010).
                                      26. Kido, Y. et al. Purification and kinetic characterization of recombinant alternative oxidase from Trypanosoma brucei brucei. Biochim.
                                          Biophys. Acta 1797, 443–450 (2010).
                                      27. Suzuki, T. et al. Alternative oxidase (AOX) genes of African trypanosomes: Phylogeny and evolution of AOX and plastid terminal
                                          oxidase families. J. Eukaryot. Microbiol. 52, 374–381 (2005).
                                      28. Murphy, A. D. & Lang-Unnasch, N. Alternative oxidase inhibitors potentiate the activity of atovaquone against Plasmodium
                                          falciparum. Antimicrob. Agents Chemother. 43, 651–654 (1999).
                                      29. Suzuki, T. et al. Direct evidence for cyanide-insensitive quinol oxidase (alternative oxidase) in apicomplexan parasite Crypto-
                                          sporidium parvum: Phylogenetic and therapeutic implications. Biochem. Biophys. Res. Commun. 313, 1044–1052 (2004).
                                      30. Sharpless, T. K. & Butow, R. A. An inducible alternate terminal oxidase in Euglena gracilis mitochondria. J. Biol. Chem. 245, 58–70
                                          (1970).
                                      31. Doussiere, J. & Vignais, P. V. AMP-dependence of the cyanide-insensitive pathway in the respiratory chain of Paramecium tetrau-
                                          relia. Biochem. J. 220, 787–794 (1984).
                                      32. Young, P. G. The SHAM-sensitive alternative oxidase in Tetrahymena pyriformis: Activity as a function of growth state and chlo-
                                          ramphenicol treatment. J. Gen. Microbiol. 129, 1357–1363 (1983).
                                      33. Coyne, R. S. et al. Comparative genomics of the pathogenic ciliate Ichthyophthirius multifiliis, its free-living relatives and a host
                                          species provide insights into adoption of parasitic lifestyle and prospects for disease control. Genome Biol. 12, R100 (2011).
                                      34. Xiong, J. et al. Genome of the facultative scuticociliatosis pathogen Pseudocohnilembus persalinus provides insight into its virulence
                                          through horizontal gene transfer. Sci. Rep. 5, 15470 (2015).
                                      35. Mallo, N., Lamas, J. & Leiro, J. M. Alternative oxidase inhibitors as antiparasitic agents against scuticociliatosis. Parasitology 141,
                                          1311–1321 (2014).
                                      36. Morais, P., Piazzon, C., Lamas, J., Mallo, N. & Leiro, J. M. Effect of resveratrol on oxygen consumption by Philasterides dicentrarchi,
                                          a scuticociliate parasite of turbot. Protist 164, 206–217 (2013).
                                      37. Iglesias, R. et al. Philasterides dicentrarchi (Ciliophora, Scuticociliatida) as the causative agent of scuticociliatosis in farmed turbot
                                          Scophthalmus maximus in Galicia (NW Spain). Dis. Aquat. Organ. 46, 47–55 (2001).
                                      38. Paramá, A. et al. Philasterides dicentrarchi (Ciliophora, Scuticociliatida): Experimental infection and possible routes of entry in
                                          farmed turbot (Scophthalmus maximus). Aquaculture 217, 73–80 (2003).
                                      39. Iglesias, R. et al. In vitro growth requeriments for the fish pathogen Philasterides dicentrarchi (Ciliophora, Scuticociliatida). Vet.
                                          Parasitol. 111, 19–30 (2003).
                                      40. Jansen, H. J. et al. Rapid de novo assembly of the European eel genome from nanopore sequencing reads. Sci. Rep. 7, 7213 (2017)
                                          ((Erratum in: Sci. Rep., 9, 7911 (2019))).

          Scientific Reports |   (2020) 10:11880 |                     https://doi.org/10.1038/s41598-020-68791-9                                                                10

Vol:.(1234567890)
www.nature.com/scientificreports/

                            41. Haas, et al. De novo transcript sequence reconstruction from RNA-seq using the Trinity platform for reference generation and
                                analysis. Nat. Protoc. 8, 1494–1512 (2013).
                            42. Folgueira, I., Lamas, J., de Felipe, A. P., Sueiro, R. A. & Leiro, J. M. Identification and molecular characterization of superoxide
                                dismutases isolated from a scuticociliate parasite: Physiological role in oxidative stress. Sci. Rep. 9, 13329 (2019).
                            43. Barua, S. & Das, B. Preparation and characterization of chitosan-based hydrogel. World J. Pharm. Pharm. Sci. 5, 2202–2208 (2016).
                            44. Iglesias, R. et al. Philasterides dicentrarchi (Ciliophora: Scuticociliatida) expresses surface immobilization antigens that probably
                                induce protective immune responses in turbot. Parasitology 126, 125–134 (2003).
                            45. Mallo, N., Lamas, J., Piazzon, C. & Leiro, J. Presence of a plant-like proton-translocating pyrophosphatase in a scuticociliate parasite
                                and its role as a possible drug target. Parasitology 142, 449–462 (2015).
                            46. Mallo, N. et al. Role of H ­ +-pyrophosphatase activity in the regulation of intracellular pH in a scuticociliate parasite of turbot:
                                Physiological effects. Exp. Parasitol. 169, 59–68 (2016).
                            47. Paramá, A., Arranz, J. A., Alvarez, M. F., Sanmartín, M. L. & Leiro, J. Ultrastructure and phylogeny of Philasterides dicentrarchi
                                (Ciliophora, Scuticociliatia) from farmed turbot in NW Spain. Parasitology 132, 555–564 (2006).
                            48. Iglesias, R., Paramá, A., Álvarez, M. F., Leiro, J. & Sanmartín, M. L. Antiprotozoals effective in vitro against the scuticociliate fish
                                pathogen Philasterides dicentrarchi. Dis. Aquat. Org. 49, 191–197 (2002).
                            49. Bustin, S. A. et al. The MIQE guidelines: minimum information for publication of quantitative real-time PCR experiments. Clin
                                Chem. 55, 611–622 (2009).
                            50. Mitchell, A. L. et al. Improving coverage, classification and access to protein sequence annotations. Nucleic Acids Res. 47, D351–
                                D360 (2019).
                            51. Käll, L., Krogh, A. & Sonnhammer, E. L. L. A combined transmembrane topology and signal peptide prediction method. J. Mol.
                                Biol. 338, 1027–1036 (2004).
                            52. Claros, M. G. & Vincens, P. Computational method to predict mitochondrially imported proteins and their targeting sequences.
                                Eur. J. Biochem. 241, 779–786 (1996).
                            53. Gasteiger, E. et al. Protein identification and analysis tools on the ExPASy server. In The Proteomics Protocols Handbook (ed. Walker,
                                J. M.) 571–607 (Humana Press, Totowa, 2005).
                            54. Gupta, R. & Brunak, S. Prediction of glycosylation across the human proteome and the correlation to protein function, pacific
                                symposium on biocomputing. Pac. Symp. Biocomput. 7, 310–322 (2002).
                            55. Wang, J., Torii, M., Liu, H., Hart, G. W. & Hu, Z. Z. dbOGAP: An integrated bioinformatics resource for protein O-GlcNAcylation.
                                BMC Bioinform. 12, 91 (2011).
                            56. Omasits, U., Ahrens, C. H., Müller, S. & Wollscheid, B. Protter: Interactive protein feature visualization and integration with
                                experimental proteomic data. Bioinformatics 30, 884–886 (2014).
                            57. Goldberg, T. et al. LocTree3 prediction of localization. Nucleic Acids Res. 42, W350-355 (2014).
                            58. Waterhouse, A. et al. SWISS-MODEL: Homology modelling of protein structures and complexes. Nucleic Acids Res. 46, W296–
                                W303 (2018).
                            59. Madeira, F. et al. The EMBL-EBI search and sequence analysis tools APIs in 2019. Nucleic Acids Res. 47, W636–W641 (2019).
                            60. Capella-Gutiérrez, S., Silla-Martínez, J. M. & Galbaldón, T. trimAl: A tool for automated alignment trimming in large-scale phy-
                                logenetic analyses. Bioinformatics 25, 1972–1973 (2009).
                            61. Jones, D. T., Taylor, W. R. & Thornton, J. M. The rapid generation of mutation data matrices from protein sequences. Comput. Appl.
                                Biosci. 8, 275–282 (1992).
                            62. Hillis, D. M. & Bull, J. J. An empirical test of bootstrapping as a method for assessing confidence in phylogenetic analysis. Syst.
                                Biol. 42, 182–192 (1993).
                            63. Kumar, S., Stecher, G., Li, M., Knyaz, C. & Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing
                                platforms. Mol. Biol. Evol. 35, 1547–1549 (2018).
                            64. Huelsenbeck, J. P. & Ronquist, F. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17, 754–755 (2001).
                            65. Meeuse, B. J. Thermogenic respiration in aroids. Annu. Rev. Plant Physiol. Plant Mol. Biol. 26, 117–126 (1975).
                            66. McDonald, A. E. & Vanlerberghe, G. C. Origins, evolutionary history, and taxonomic distribution of alternative oxidase and plas-
                                toquinol terminal oxidase. Physiol. Part D Genom. Proteomics 1, 357–364 (2006).
                            67. Rodríguez-Armenta, C. et al. Alternative mitochondrial respiratory chains from two crustaceans: Artemia franciscana nauplii and
                                the white shrimp, Litopenaeus vannamei. J. Bioenergy Biomembr. 50, 143–152 (2018).
                            68. Dinant, M., Baurain, D., Coosemans, N., Joris, B. & Matagne, R. F. Characterization of two genes encoding the mitochondrial
                                alternative oxidase in Chlamydomonas reinhardtii. Curr. Genet. 39, 101–108 (2001).
                            69. Roberts, C. W. et al. Evidence for mitochondrial-derived alternative oxidase in the apicomplexan parasite Cryptosporidium parvum:
                                A potential anti-microbial agent target. Int. J. Parasitol. 34, 297–308 (2004).
                            70. Polidoros, A. N., Mylona, P. V. & Arnholdt-Schmitt, B. Aox gene structure, transcript variation and expression in plants. Physiol.
                                Plant. 137, 342–353 (2009).
                            71. Costa, J. H. et al. Stress-induced co-expression of two alternative oxidase (VuAox1 and 2b) genes in Vigna unguiculata. J. Plant
                                Physiol. 167, 561–570 (2010).
                            72. Umbach, A. L. & Siedow, J. N. The cyanide-resistant alternative oxidases from the fungi Pichia stipitis and Neurospora crassa are
                                monomeric and lack regulatory features of the plant enzyme. Arch. Biochem. Biophys. 378, 234–245 (2000).
                            73. Siedow, J. N., Umbach, A. L. & Moore, A. L. The active site of the cyanide-resistant oxidase from plant mitochondria contains a
                                binuclear iron center. FEBS Lett. 362, 10–14 (1995).
                            74. Anderson, M. E. & Nordlund, P. A revised model of the active site of alternative oxidase. FEBS Lett. 449, 17–22 (1999).
                            75. Berthold, D. A., Andersson, M. E. & Nordlund, P. New insight into the structure and function of the alternative oxidase. Biochim.
                                Biophys. Acta 1460, 241–254 (2000).
                            76. Young, L. et al. The alternative oxidases: Simple oxidoreductase proteins with complex functions. Biochem. Soc. Trans. 41, 1305–
                                1311 (2013).
                            77. Bonnefoy, N., Fiumera, H. L., Dujardin, G. & Fox, T. D. Roles of Oxa1-related inner-membrane translocases in assembly of respira-
                                tory chain complexes. Biochem. Biophys. Acta 1793, 60–70 (2009).
                            78. Allen, K. N., Entova, S., Ray, L. C. & Imperiali, B. Monotopic membrane proteins join the fold. Trends Biochem. Sci. 44, 7–20 (2019).
                            79. Umbach, A. L. & Siedow, J. N. Covalent and noncovalent dimers of the cyanide-resistant alternative oxidase protein in higher plant
                                mitochondria and their relationship to enzyme activity. Plant Physiol. 103, 845–854 (1993).
                            80. Berthold, D. A. & Stenmark, P. Membrane-bound di-iron carboxylate proteins. Ann. Rev. Plant Biol. 54, 497–517 (2003).
                            81. Shiba, T. et al. Structure of the Trypanosome cyanide-insensitive alternative oxidase. Proc. Natl. Acad. Sci. U.S.A. 110, 4580–4585
                                (2013).
                            82. Moore, A. L., Umbach, A. L. & Siedow, J. N. Structure-function relationships of the alternative oxidase of plant mitochondria: A
                                model of the active site. J. Bioenergy Biomembr. 27, 367–377 (1995).
                            83. Pennisi, R. et al. Molecular evolution of alternative oxidase proteins: A phylogenetic and structure modelling approach. J. Mol.
                                Evol. 82, 207–218 (2016).
                            84. De Felipe, A. P., Lamas, J., Sueiro, R. A., Folgueira, I. & Leiro, J. M. New data on flatfish scuticociliatosis reveal that Miamiensis
                                avidus and Philasterides dicentrarchi are different species. Parasitology 29, 1–18 (2017).

Scientific Reports |   (2020) 10:11880 |                    https://doi.org/10.1038/s41598-020-68791-9                                                               11

                                                                                                                                                               Vol.:(0123456789)
You can also read